You are on page 1of 13

International Journal of Heat and Mass Transfer 173 (2021) 121248

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/hmt

Review

Laminar natural convection in a square cavity with 3D random


roughness elements considering the compressibility of the fluid
Boqi Ren a, Chung-Gang Li a,b,∗, Makoto Tsubokura a,b
a
Graduate School of System Informatics, Kobe University, 1-1 Rokkodai, Nada-ku, Kobe, Hyogo 657 8501, Japan
b
Complex Phenomena Unified Simulation Research Team, RIKEN, Advanced Institute for Computational Science, 7-1-26 Minatojima-minami-machi, Chuo-ku,
Kobe, Hyogo 650-0047, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Natural convection in a three-dimensional square cavity with random artificial roughness on both vertical
Received 26 October 2020 walls is studied numerically for a Rayleigh number of 106 . Based on consideration of realistic conditions,
Revised 9 March 2021
the roughness is generated using a given power spectrum density, and a compressible solver with precon-
Accepted 18 March 2021
ditioning that uses a dual time-stepping method is established to handle the low velocity of the natural
Available online 5 April 2021
convection flow. The compressible solver demonstrates satisfactory accuracy in terms of the average Nus-
Keywords: selt number when compared with incompressible benchmark solutions. However, the results for the local
Natural convection Nusselt number from the compressible and incompressible solvers differ. As the temperature difference
3D random roughness between the two vertical sidewalls increases, the maximum value of the local Nusselt number increases,
Compressible flow but in the downstream region, the local Nusselt number actually decreases. The cavity with rough side-
Nusselt number walls has a similar Nusselt number to the average Nusselt number of a hot sidewall in the case of a
smooth cavity. Because of the effects of the roughness on the thermal and velocity boundary layers, the
rough cavity case shows a better heat transfer performance than the smooth cavity case in the upstream
region, but in the downstream region, the heat transfer in the rough case becomes worse than that in
the smooth case.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction to use a compressible flow solver because of the conspicuous den-


sity variations of fluids. However, when compared with the speed
In modern industry, many facilities operate under conditions of sound, the magnitude of the fluid speed is too small to allow a
that involve high temperature differences, e.g., nuclear reactor compressible solver to be used directly. For this reason, Weiss and
cooling, and the cooling systems used in construction and solar Smith [8] adopted a preconditioning method for application to nat-
collectors [1–3], so, in that case, the influence of compressibility ural convection in a horizontal annulus and their results showed
of fluids on the heat transfer performance becomes conspicuous. that the accuracy of this method was satisfactory under the condi-
On the other hand, to enhance the heat transfer, the roughness or tion where the Rayleigh number was equal to 4.7 × 104 . Yamamoto
the presence of obstacles on the surfaces from which heat trans- et al. [9] simulated natural convection in a horizontal circular pipe
fer takes place has been researched widely, but most studies only using the same method to deal with compressible flows at low ve-
consider incompressible fluids [4–6]. Therefore, a study of the heat locities and proved that the preconditioning method could achieve
transfer considering both roughness and the compressibility of flu- good agreement with the experimental results in terms of the ther-
ids is necessary and significant. mal boundary layers and the Nusselt number under the conditions
For the case of a high temperature difference (i.e., >30 K), Gray of a temperature difference of 32.5 K and the Rayleigh number
and Giorgini [7] recommended considering the compressibility of Ra = 105 . In addition, the preconditioning method was shown to
the fluid during numerical calculations, thus making it necessary accelerate the simulation toward convergence under steady-state
conditions [10].
Buoyancy driven flows in square cavities have previously been

Corresponding author at: Computational Fluid Dynamics Laboratory, Depart- studied widely. Bajorek and Lloyd [11] used experimental methods
ment of Computational Science, Graduate School of System Informatics, Kobe Uni- to investigate the effects of rectangular partitions on the horizon-
versity, 1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan. tal walls in a square cavity on the heat transfer between the two
E-mail addresses: bochiren1992@stu.kobe-u.ac.jp (B. Ren), vertical walls. Their results showed that, when compared with a
cgli@aquamarine.kobe-u.ac.jp (C.-G. Li), tsubo@tiger.kobe-u.ac.jp (M. Tsubokura).

https://doi.org/10.1016/j.ijheatmasstransfer.2021.121248
0017-9310/© 2021 Elsevier Ltd. All rights reserved.
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

ber was 106 . However, there was an increment in the surface area
Nomenclature of 32% when using the single roughness element in comparison
with the smooth wall case. The mean Nusselt number of the case
L1 , L2 , L3 width, height and depth of the square cavity where two roughness elements were present showed a slight in-
A area of sidewalls crease in the mean Nusselt number (3.88% higher) when compared
C power spectral density (PSD) with the single roughness element case.
d distance from the cell center to the closest geom- In addition to the rectangular roughness elements, sinusoidal
etry element roughness elements have also been studied recently by many re-
g gravitational acceleration searchers. Bhardwaj et al. [13] simulated the flow field in a porous
h height of individual roughness element right-angled triangular cavity with a sine-wave-shaped left wall
k theermal conductivity and a smooth bottom wall to determine the effects of the fluctu-
Nu Nusselt number ating hot wall on the heat transfer and entropy generation prop-
P pressure erties of the cavity. Their results showed that, under the laminar
Pr Prandtl number flow condition, the undulation of the hot wall caused the strength
q wavevector of the vortex to be enhanced; in addition, both the average Nus-
R gas constant selt number and the maximum Nusselt number were higher than
Ra Rayleigh number based on cavity in the zero-undulation case. The increment in the maximum Nus-
height, gβ0 T L2 3 P r/ν0 2 selt number was 53% for a Rayleigh number of 106 and a Darcy
Rq root mean square roughness number of 0.01. However, in contrast, Yousaf and Usman [14] ob-
T temperature tained a lower average Nusselt number along a hot wall with si-
t physical time nusoidal roughness elements. At an Ra of 106 and using 10 rough-
u1 , u2 , u3 velocity components in x1 , x2 and x3 directions ness elements, the average Nusselt number decreased by 12.25%
x1 , x2 , x3 spatial coordinates when compared with that in the smooth wall case. Shahriari et al.
[15] also observed a reduction in the Nusselt number. At Ra = 105 ,
Greek symbols
with eight sinusoidal roughness elements on both hot and cold
β thermal expansion coefficient
walls in a square cavity, the Nusselt number decreased by approx-
T temperature difference, (TH − TC )
imately 20% when compared with that of the smooth wall case.
x 1 , x 2 , x 3 mesh size in x1 , x2 and x3 directions
In most previous experimental or numerical researches, sim-
t physical time step
ple geometric shapes, such as thin sheets and rectangular or si-
τ artificial time step
nusoidal shapes, were selected as the shapes of roughness ele-
δi j Kronecker delta
ments [12,14,15]. In contrast, to represent the roughness existing in
δx i central difference operator
nature, a three-dimensional artificial random roughness has been
γ adiabatic index
generated in the present study. The roughness elements are ran-
 preconditioning matrix
dom in shapes and heterogeneous in sizes. Furthermore, by tak-
θ nondimensional temperature,
ing the working conditions of practical industrial applications into
(T − Tc )/(Th − Tc )
account, a compressible solver with an all-speed Roe scheme has
ρ density
been established as the appropriate numerical method to investi-
τ artificial time
gate buoyancy-driven flows.
μ dynamic viscosity
ν kinematic viscosity 2. Physical model and governing equations
φ arbitrary physical value

Subscript Because the three-dimensional geometries of random rough


C cold wall surfaces cannot be ignored in this work, the numerical simulations
H hot wall of the flows in the square cavity are performed under 3D condi-
IC interface cell tions. Fig. 1 shows the physical model of current study with dif-
IP image point ferentially heated vertical sidewalls and adiabatic top and bottom
l local value surfaces. A high temperature of TH and a low temperature of TC ,
overall average value over an entire sidewall which are both constant and isothermal, are set on the left and
p primitive form of variables right sidewalls, respectively. The no-slip boundary condition is im-
w wall posed on the top, bottom and both sidewalls of the cavities, but
0 ambient or initial value slip boundary condition is imposed on both x3 -direction bound-
aries (at x3 /L3 = 0 and x3 /L3 = 1 planes). For all cavity flow, air
Superscript with a Prandtl number (Pr) of 0.72 is selected as the working fluid.
∗ nondimensional value The initial temperature of the air contained in the core of the
k artificial time square cavity is T0 = (TH + TC )/2 and the pressure is 101,300 Pa.
n physical time The artificial random surface roughness is generated based on
a given power spectrum density [16,17], as illustrated in Fig. 2. In
this research, the root mean square roughness
 (Rq) is equal to 1.5%
1  2
n
smooth cavity, the presence of the partitions caused the heat trans-
of the cavity width (L1 ), where Rq = n hi , and hi is the height
fer to be reduced by 12% and 21% at high Grashof numbers and low i=1
Grashof numbers, respectively. Shakerin et al. [12] used experimen- of each roughness element from the mean plane of the roughness.
tal methods to study the heat transfer from hot walls with differ- On the mean plane’s position, the heights of peaks and depths of
ent numbers of rectangular roughness elements within a square valleys all add to zero. Because of the random distribution of the
enclosure. For a single roughness element, the mean Nusselt num- roughness elements, the depth-to-width aspect ratio (L3 /L1 ) is se-
ber along the hot wall showed an increase of 12% when compared lected to be 2 and the arrangement of the roughness elements is
with the smooth wall under the condition that the Rayleigh num- symmetrical with respect to the middle of the depth of the cav-

2
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

⎡ ⎤
ρUi
⎢ ρ ui u1 + pδi1 − μAi1 ⎥
⎢ ρ u u + pδ − μA ⎥
Fi = ⎢ i 2 i2 i2 ⎥, ∀i, j = 1, 2, 3, (3)
⎣ ρ ui u3 + pδi3 − μAi3 ⎦
(ρ e + p)ui − μAi j u j − k ∂∂xTi
and
⎡ 0

⎢ 0 ⎥
S = ⎢ −(ρ − ρ0 )g ⎥, (4)
⎣ ⎦
0
−(ρ − ρ0 )gu2
where Ai j = ∂ u j /∂ xi + ∂ ui /∂ x j − 2/3(∇ · u )δi j and δi j is the Kro-
necker delta. The relationship between the pressure and the den-
sity of the air is represented by the ideal gas equation, where
P = ρ RT . According to Sutherland’s Law [18], the viscosity (μ) and
the thermal conductivity (k) of the working fluid are determined
as follows:

T 32 T + 110.4
μ ( T ) = μ0 0
, (5)
T0 T + 110.4
Fig. 1. Simulation settings of the square cavity.
μ(T )γ R
k (T ) = , (6)
(γ − 1 )Pr
ity. As noted above, in many previous works, the existence of the
where ρ0 = 1.18 kg/m3 , g = 9.81 m/s2 , μ0 = 1.85 × 10−5 N · s/m2 ,
roughness will reduce the effective volume of fluid in the cavity.
T0 = 298.06 K, γ = 1.4, R = 287 J/kg, and P r = 0.72.
Therefore, to maintain a constant value for the effective volume,
rough surfaces have been installed on both hot and cold sidewalls
3. Numerical method
in this research, and the rough surface on the cold wall is gener-
ated by direct translation of the rough surface on the hot wall.
3.1. Compressible solver
To deal with the natural convection that occurs because of the
buoyancy force, the following compressible governing equation has
In natural convection, the magnitude of the fluid velocity due to
been considered:
the buoyancy force is generally smaller than the speed of sound by
∂ U ∂ F1 ∂ F2 ∂ F3 several orders of magnitude. For this reason, the original compress-
+ + + = S. (1) ible solvers are unsuited to analysis of this type of flow field. To
∂ t ∂ x1 ∂ x2 ∂ x3
increase the efficiency for solution of natural convection problems
U and Fi can be written in following format: involving compressible flows, a preconditioning method is used to
resolve the governing equations, which is developed by Weiss and
⎡ ρ ⎤ Smith [8]. When combined with the Roe scheme [19] and the dual
⎢ρ u 1 ⎥ time-stepping method [20], the new governing equation is shown
U = ⎢ρ u 2 ⎥, (2) as follows:
⎣ ⎦
ρ u3 ∂ Up ∂ U ∂ F1 ∂ F2 ∂ F3
ρe  + + + + = S, (7)
∂τ ∂ t ∂ x1 ∂ x2 ∂ x3

Fig. 2. 3D random artificial roughness and the power spectrum.

3
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

where  is the preconditioning matrix provided by Weiss and


Smith, U p is the primitive form of the variables [P, u1 , u2 , u3 , T ],
and U is the conservative form of [ρ , ρ u1 , ρ u2 , ρ u3 , ρ e]. τ and t
represent the artificial and physical times, respectively.
The discretized form of Eq. (7) is

U pk+1 − U pk3U k+1 − 4U n + U n−1 1




 + + F k+1 − F k+1
τ 2t x1 1i+1/2, j,k 1i−1/2, j,k
1

1


+ F2k+1 − F2k+1 + F3k+1 − F3k+1 = Sk . (8)
x2 i, j+1/2,k i, j−1/2,k x3 i, j,k+1/2 i, j,k−1/2

The iterations of the artificial time and the physical time are
indicated by the superscripts k and n shown in Eq. (8). It should
be noted that when the iteration of the artificial time reaches con-
vergence, the residual of ∂ U p /∂τ is less than 10−3 , and the mag-
nitude of the artificial term at the (k+1)th step is equivalent to
the magnitude of the physical time term at the (n+1)th step, then
Eq. (8) would then return approximately to the original Navier-
Stokes equation.
Eq. (8) can be expressed in the following format as:
I 3 
+  −1 M +  −1 δx1 Akp + δx2 Bkp + δx3 C pk U p =  −1 Rk ,
τ 2t
(9)
where M = ∂ U/∂ U p , A p = ∂ F1k /∂ U p , = − Rk Sk
Fig. 3. Configuration of immersed boundary method.
(3U − 4U + U )/2t − (δx1 F1 + δx2 F2k + δx3 F3k ),
k n n −1 k and δxi is
the central difference operator. Xu et al. [21] recommended setting
the pseudo-time step, τ , to infinity then the Newton lineariza- 3.2. Immersed boundary method for compressible flows
tion error will be zero in each physical time step. Eventually,
Eq. (9) can be rewritten into Eq. (10): Because of the complex geometry of the random rough surface,
3  the immersed boundary method(IBM) has been applied [23] to en-
 −1 M +  −1 δx1 Akp + δx2 Bkp + δx3 C pk U p =  −1 Rk , (10) sure that the geometrical details are reproduced as accurately as
2t
possible. The key point of IBM is to simulate the existence of ge-
The lower-upper symmetric-Gauss-Seidel (LUSGS) implicit ometry by imposing the physical values into the numerical grids in
method is then used to solve Eq. (10). the vicinity of geometry. Fig. 3 shows the configuration of IBM. The
It is helpful to divide the flux term given in Eq. (3) into an in- cells which are defined as interface cell (IC) has the characteristic
viscid term and a viscid term, as shown in Eqs. (11) and (12), re- that the distance (d) from the cell center to the closest geometry
spectively, to perform the calculation of Rk in Eq. (10). The inviscid element is less than one cell size. Image point (IP) is the point that
term in Eq. (11) can then be discretized into Eq. (13) using the Roe the distance from which to IC is equal to d in the normal direction
scheme. of the geometry element. IC and IP must be on the same side of
⎛ ρ ui ⎞ geometry. Then, according to the relative locations of surrounding
⎜ρ ui u1 + pδi1 ⎟ cells to IP, the physical value at IP, φIP , can be calculated by bilinear
Finviscid = ⎜ρ ui u2 + pδi2 ⎟, (11) interpolation:
⎝ ⎠
ρ ui u3 + pδi3 φIP = w1 φ1 + w2 φ2 + w3 φ3 + wIC φIC , (16)
(ρ e + p)ui
where w1 , w2 , w3 and wIP can be obtained using VanderMonde
⎛ ⎞ matrix [23]. Tullio et al.[24] recommended applying linear interpo-
0
⎜ μA i 1 ⎟ lation to evaluate φIC , shown in the following equations:
⎜ μA i 2 ⎟ 
Fviscous = −⎜ ⎟, (12) ∂φ φIP − φw
⎝ μA i 3 ⎠ φw = φIC − d = φIC − d, (17)
∂n 2d
μAij u j + k ∂∂xT i

1
Finviscid,i+1/2
1
= [FR (U ) + FL (U )] + Fd , (13)
φIC = (φw + φIP ), (18)
2
2
In present study, the no-slip and isothermal conditions are im-
where Fd is the Roe dissipation term.
plemented on the walls, so φw is an assigned value φassign , thus
For the two convective terms (FR and FL ) shown in Eq. (13), the
the Dirichlet condition should be considered. Then according to
values at the cell interfaces are calculated by interpolation of the
Eqs. (16)–(18), φIC can be presented as follows:
corresponding values from a group of cell centers through a fifth-
order MUSCL approach [33] without limiter functions, which are w1 φ1 + w2 φ2 + w3 φ3 + φassign
shown in the following way: φIC = Dirichlet condition, (19)
2 − wIC
1 However, for pressure, Neumann condition, which considers
UL1+1/2 = [−3Ui+2 + 27Ui+1 + 47Ui − 13Ui−1 + 2Ui−2], (14)
60 φw = φIP , should be used, then φIC can be presented as follows:

1 w 1 φ1 + w 2 φ2 + w 3 φ3
UR1+1/2 = [2Ui+3 − 13Ui+2 + 47Ui+1 + 27Ui − 3Ui−1 ]. (15) φIC = Neumann condition, (20)
60 1 − wIC + ε

4
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 4. Numerical grids distributions of smooth (a) and rough (b) cases at x3 = 0.5 plane.

where ε is a small value to prevent the denominator becoming Table 1


Convergence behaviors of hydrodynamic quantities and Nusselt numbers.
zero. Because when IC is very close to the wall, wIC will approach
1. Ra Ref. Numerical method u2 ∗ max (x1 /L1 ) Nu
In authors’ previous study [25], the natural convection of a 10 6
E Tric [27] Incompressible 0.258 (0.033) 8.877
heated sphere has been simulated through the IBM mentioned A Xu [28] Incompressible 0.258 (0.037) 8.881
above. The comparison of the averaged Nu shows good agreement Present Compressible (10 K) 0.260 (0.039) 8.854
with the result of Jia and Gogos [26], and the error is only 0.343%. Present Compressible (50 K) 0.259 (0.041) 8.874
Present Compressible (120 K) 0.264 (0.043) 8.787

3.3. Building cube method


numbers can be obtained using Eqs. (21) and (22). The average
The building cube method (BCM) is used to provide higher
Nusselt number over entire hot sidewall is calculated through
computation efficiency and better load balancing on a massively
Eq. (23). The maximum nondimensional vertical velocity at the
parallel computing system [25]. The BCM divides the calculation
center of the hot sidewall across the nondimensional width of the
domain into several cubes, and each cube contents 163 meshes. In
cavity (x1 /L1 ) is also presented
 in the table. The nondimensional
any individual cube, the mesh spacings along the three orthogonal
velocity is defined as u∗i = ui / gβ0 T L2 .
axes (directions x1 , x2 , and x3 ) are same, which allows the details
of the flow fields between two roughness elements to be captured 1 L2
Nu = ∫ N ul d x2 , (21)
more easily. L2 0

L2 ∂T
4. Results and discussion N ul = k (T ) , (22)
k0 (Th − Tc ) ∂ x1
4.1. Investigation of fluid compressibility in enclosed square cavity  A
1
Nuoverall = N ul dA, (23)
To discuss the effects of the compressible fluid on heat trans- A 0

fer, 3D cavities, shown in Fig. 1, with smooth hot and cold where k0 = k(T0 ) and A = L2 · L3 .
sidewalls are investigated in this section. For all cavities, the The average Nusselt numbers of the three cases with various
Rayleigh number remains constant, where the Rayleigh number temperature differences from Table 1 show good agreement with
Ra = gβ0 T L2 3 P r/ν0 2 = 106 and all fluid properties are determined the results from previous studies at the Ra of 106 . The maximum
based on the ambient temperature. deviation is 1.06%, which is the error between the case with the
With regard to the influence of the compressibility of the work- temperature difference of 120 K and the results of A Xu [28].
ing fluid on the heat transfer, three different temperature differ- Fig. 5 presents the local Nusselt number distributions at the mid-
ences (10 K, 50 K, and 120 K) have been studied and the mesh dle depth on the hot sidewall. For the compressible solver, it is
distributions have been shown in Fig. 4(a). The finest grids are im- evident that as T increases, the pattern of the local Nu numbers
plemented in the vicinity of two sidewalls, and to save the compu- on the hot wall becomes steeper, i.e., the maximum value of the
tational resources, the coarser grids are implemented in the core local Nu increases, but the cases with larger T values will have
regions of the cavities. There are 256 finest grids along with the smaller local Nu magnitudes in the x2 /L2 > 0.3 region. Between the
height (L2 ) direction of the enclosed cavities, and the distributions case with the temperature difference of 120 K and the results of
of grids along the depth (L3 ) direction are uniform for all three Kuyper et al. [29], the maximum local Nusselt numbers reach their
cases. The Courant number (CFL) for all cases is set as 0.5 where largest deviation of 6.233%. For the vertical velocity, Table 1 indi-
CF L = max(ui · t/xi ), i = 1, 2, 3 to represent the components cates that as the temperature difference increases, the location of
acting along the three axes in the coordinate system. the maximum vertical velocity moves away from the hot sidewalls.
Table 1 shows the average Nusselt numbers across the x2 - This phenomenon where the distribution of the local Nusselt num-
direction and the maximum values of the local Nusselt numbers ber changes with the temperature difference was also reported in
on the hot sidewalls across the height at x3 /L3 = 0.5 plane; these [30] and it was attributed to the temperature dependences of the

5
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 5. Local Nusselt numbers of smooth cases with various temperature differences.

Table 2 Table 3
Values of average Nu for grid sensitivity study. Maximum values of the local Nusselt numbers and
the average Nusselt numbers for the smooth and
Mesh resolution Nuoverall Difference (%) rough cases.
512 8.719 -
Smooth Rough Difference (%)
256 8.717 0.023
128 7.765 10.942 Nulmax 18.134 60.207 +232.012
Nuoverall 8.874 8.717 -1.769

fluid properties, as shown in Eqs. (5) and (6). In summary, a tem-


perature difference between two heated sidewalls will change the grid 128 shows a different profile of temperature from the other
local Nusselt number distribution but will have no effect on the two mesh resolutions. According to the arrangement of roughness
average Nusselt number for a Rayleigh number of 106 . Therefore, elements in current study, for the resolution of grid 256, there are
it becomes necessary to consider adopting a compressible solver at least 4 grids for one peak or one valley on the rough surface,
when local region heat transfer is the research target. but for the resolution of grid 128, this number is only 2 for one
peak or one valley.
4.2. Grid sensitivity for the enclosure with rough sidewalls Thus, all subsequent results in the following section are based
on the mesh resolution of grid 256.
Due to the complex 3D random roughness, the Immersed-
boundary method which is mentioned in Section 3.2 requires suit- 4.3. Influence of roughness on heat transfer
able girds to capture the geometric features of the roughness.
Meanwhile, for the simulation of the fluid among the roughness Fig. 7 shows the local Nusselt number distributions on the
elements, an appropriate mesh resolution is necessary. Therefore, hot sidewalls of the smooth and rough cavity cases. For both the
in this part, the cavity with rough sidewalls is filled by numeri- smooth case and the rough case, the local Nusselt numbers show
cal grids with 3 different resolutions shown in Table 2. The mesh a trend of being higher in the upstream region and lower in the
resolutions of 128 and 512 are respectively generated by coarsen- downstream region. Because of the existence of the peaks and val-
ing or refining the grids based on the resolution of 256, shown leys on the sidewalls, the local Nusselt number distribution on the
in Fig. 4(b). For convenience, we use grid 128, grid 256 and grid rough surface shows good agreement with the distribution of the
512 to represent these three mesh resolutions. Table 2 shows the roughness elements and the local Nusselt numbers on the peaks
average Nusselt numbers over the entire hot sidewall from these are larger than those in the valleys. Table 3 shows the average Nus-
3 different grid strategies. After a comparison between the results selt numbers over the entire hot sidewall and the maximum values
of grid 256 and grid 512, the average Nusselt numbers are almost of the local Nusselt numbers for the two cases, where the symbols
same and the difference is only 0.023%. However, the difference of + and – represent that compared with the results of the smooth
between grid 512 and grid 128 is 10.942%. case, the increase or decrease of the investigated values due to the
Further, Fig. 6 shows the nondimensional temperature distri- rough sidewalls respectively. In Table 3, the largest local Nusselt
butions across the horizontal centerline of the cavities’ midplane number in the rough cavity case is much larger than that in the
(x2 /L2 = 0.5 line at x3 /L3 = 0.5 plane). The profiles show that the smooth case; however, in contrast, the average Nusselt number of
temperature distributions of grid 256 and gird 512 almost over- the rough case is slightly smaller than that in the smooth case by
lapped near the hot sidewall, but due to the insufficient resolution, 1.769%.

6
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 6. Nondimensional temperature profiles.

Fig. 7. Distributions of the local Nu values of the (a) smooth and (b) rough cases on the hot walls.

Table 4 When compared with the smooth case, the local average Nus-
Local average Nusselt numbers in the upstream and downstream regions.
selt number in the upstream region in the rough case is higher,
Smooth Rough Difference (%) but in the downstream region, the local average Nusselt number
Upstream (x2 /L2 = 0 to 0.1) 17.613 18.622 +5.729 in the rough case is smaller than that in the smooth case. To ex-
Downstream (x2 /L2 = 0.75 to 1) 2.822 2.743 -2.799 plain this phenomenon in greater depth, the temperature distribu-
tion on a slice at a quarter of the depth of the cavity is shown in
Fig. 8, where the nondimensional temperature is calculated using
Table 4 presents the local average Nusselt numbers of the two the following equation: θ = (T − Tc )/(Th − Tc ). In the upstream re-
cases above in the upstream and downstream regions. Here, the gion, the peaks will reduce the thermal boundary layer thickness,
local average Nusselt number is considered to be the integral of the while the valleys can make the thermal boundary layer thicker,
Nusselt number on a local part of the heated sidewall, as defined which means that the thermal boundary layer will undergo drastic
in Eq. (24): changes in thickness within this region. However, as the thermal
boundary layer develops, the thickness will inevitably increase, and
1 Al then the effects of the roughness elements on the thermal bound-
Nul = ∫ N ul dA, (24)
Al 0 ary layer will decrease. From Fig. 8(c) and (d), the isothermal sur-
faces are distorted by the roughness elements near the rough side-
where the subscript l indicates a specific local part of the heated
sidewall.

7
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 8. Temperature distributions of smooth and rough cases: (a, b) slices at x3 /L3 =0.25 plane (c, d) isothermal surface in half cavity from x3 /L3 =0 to 0.5.

wall, but in the core region of the cavities, the temperature strati- in the upstream region, while in the downstream region, the local
fications are almost identical for smooth and rough cases. average Nusselt number is smaller than that in the smooth case.
Fig. 9(a) and (b) show the nondimensional velocity magni- Fig. 10 shows the contours of the nondimensional velocity mag-
tudes of the two cases on three different planes (x1 /L1 = 0.005, nitude of both smooth and rough cases near the hot sidewall at
x2 /L2 = 0.5 and x3 /L3 = 0.25) respectively. From this figure, when four different x1 -x2 planes (x3 /L3 = 0.125, 0.25, 0.375 and 0.5),
the flow approaches the peaks on the rough surface, there is a where the black curves are the profiles of hot sidewalls at these
slight acceleration in the magnitude of the velocity when com- planes. From the figure, the velocity magnitudes of the flow field in
pared with the smooth case, but after the peak has been crossed, the vicinity of the rough wall are smaller than smooth case gener-
the ensuing valley will cause the fluid velocity to decelerate ally, and only part of the fluid will be accelerated due to the rough-
sharply. Unlike the influence of the rough surface on the thermal ness. In the region where the shape of roughness elements does
boundary layer, the influence of the rough surface on the magni- not change drastically, the roughness plays a role more in slowing
tude of the velocity in the downstream region is obvious. This can down the velocity of adjacent fluid rather than accelerating it.
therefore be considered to be one explanation as to why the local Fig. 11(a) and (b) show the velocity vector fields in the half re-
Nusselt number distribution presents the tendency where, in the gion located close to the hot sidewall at the same location used in
case with the rough heated sidewalls, the local Nusselt number is Fig. 10(a) and (b). The figure containing the velocity vectors shows
higher than that in the smooth case in the vicinity of the peaks no evident fluid circulation to be found in the area between the
two roughness elements on the x1 -x2 plane at Ra = 106 . In addi-

8
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 9. Nondimensional velocity magnitudes (where (a) and (b) show the smooth and rough cases, respectively) and vectors (where (c) and (d) show the smooth and rough
cases, respectively).

Fig. 10. Contours of Nondimensional velocity magnitude of smooth (a) and rough (b) cases at different x1 -x2 planes.

9
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 11. Velocity vector fields of smooth (a) and rough (b) cases at different x1 -x2 planes.

Fig. 12. Nondimensional temperature gradients across the height of the cavity at two different measurement locations (black curves: profiles of the rough surface at the two
measurement locations).

tion, the eddies that occur near the core regions of the two cavities tions. These two locations are two different lines parallel to the x2 -
do not show particularly notable differences. direction(x3 /L3 = 0.55 line and x3 /L3 = 0.8 line) at x1 /L1 = 0.05
In Fig. 12, the components of the nondimensional temperature plane starting from the mean plane of the hot rough sidewall. The
gradients in the x1 -direction have been measured in the cavities black curves shown in Fig. 12 are the profiles of the hot rough
with both rough and smooth sidewalls at two measurement loca- surfaces at x1 -x2 planes where two measurement locations lie re-

10
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 13. Nondimensional temperatures at the center of the adiabatic surface across the cavity height.

Fig. 14. Magnitudes of the velocity component in the x3 direction.

spectively. The figure shows that in the upstream regions, the tem- mean plane of the rough surface. In this figure, the changes in the
perature gradient fluctuates dramatically and this trend follows colors near the roughness elements and the speed vectors demon-
the distribution of the roughness elements perfectly, which means strate that the geometry of the 3D random roughness structure
that higher peaks will bring increased heat transfer. However, this gives the fluid the possibility of bypassing the roughness elements
dominance of the roughness decreases in the end region, partic- and no evident eddies were found. However, in two-dimensional
ularly in the downstream region, and after the nondimensional cavity flow with homogeneous roughness elements on the cavity
height exceeds 0.8, the patterns from the two cases almost over- sidewalls [15,16], the fluid will become trapped more easily be-
lap. In Fig. 13, at the center of the adiabatic surface and across tween two adjacent roughness elements, and this represents a ma-
the heights of the cavities, the temperatures in both the rough jor cause of the reduction in the heat transfer.
and smooth cases have been measured. The temperature curves
show very good agreement, thus indicating that the random ar- 5. Conclusions
tificial roughness cannot influence the heat transfer that occurs in
the core region of the square cavity. Natural convection in a cavity with 3D random roughness ele-
Figs. 14 and 15 show the effects of the surface roughness on ments on its sidewalls has been investigated at a Rayleigh number
the velocity component in the x3 -direction. The slices in Fig. 14 are of 106 using a compressible solver. The roughness was generated
located at different x1 -x2 planes. The magnitudes of this veloc- using a given power spectrum density and the temperature differ-
ity component shows a trend where the initial magnitude is very ence between the two sidewalls was 50 K. The distribution of the
small in the upstream region and then increases to reach a max- local Nusselt number on the hot sidewall showed that cavities with
imum in the middle region; however, in the downstream region, both rough and smooth sidewalls followed a trend where the local
deceleration occurs. Furthermore, the effect of the rough surface Nusselt number was higher in the upstream region and lower in
cannot reach the core region of the cavity. The slice shown in the downstream region. The maximum value of the local Nusselt
Fig. 15 is located at a position of 0.5% of the cavity width from the number for the rough surface is 2.32 times larger than that in the

11
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

Fig. 15. Flow field among the roughness elements of the hot sidewall.

smooth case, but the average Nusselt number is 1.8% smaller than [4] X. Zhu, R.J.A.M. Stevens, R. Verzicco, et al., Roughness-facilitated local 1/2 scal-
that in the smooth case. The peaks on the rough surface will re- ing does not imply the onset of the ultimate regime of thermal convection,
Phys. Rev. Lett. 119 (15) (2017) 154501.
duce the thermal boundary layer thickness but the corresponding [5] S. Wagner, O. Shishkina, Heat flux enhancement by regular surface roughness
valleys will increase this thickness in the upstream region. How- in turbulent thermal convection, J. Fluid Mech. 763 (2015) 109.
ever, in the downstream region, the thermal boundary layer thick- [6] B.N. Prasad, J.S. Saini, Effect of artificial roughness on heat transfer and friction
factor in a solar air heater, Sol. Energy 41 (6) (1988) 555–560.
ness is thick enough to ensure that this effect is no longer obvious. [7] D.D. Gray, A. Giorgini, The validity of the Boussinesq approximation for liquids
Near the hot sidewall, the peaks accelerate the flow field slightly, and gases, Int. J. Heat Mass Transf. 19 (5) (1976) 545–551.
but the valleys cause a conspicuous deceleration of the flow field; [8] J.M. Weiss, W.A. Smith, Preconditioning applied to variable and constant den-
sity flows, AIAA J. 33 (11) (1995) 2050–2057.
this phenomenon exists in both the upstream and downstream re-
[9] S. Yamamoto, D. Niiyam, B.R. Shin, A numerical method for natural convection
gions. Therefore, the Nusselt number on the rough surface in the and heat conduction around and in a horizontal circular pipe, Int. J. Heat Mass
upstream region shows a larger average value than that of the Transf. 47 (26) (2004) 5781–5792.
[10] E. Turkel, Preconditioning techniques in computational fluid dynamics, Annu.
smooth case in the same region, but will show a smaller average
Rev. Fluid Mech. 31 (1) (1999) 385–416.
value in the downstream region. However, the fluid will bypass the [11] S.M. Bajorek, J.R. Lloyd, Experimental investigation of natural convection in
elements in the 3D simulations, rather than become trapped be- partitioned enclosures, J. Heat Transf. 104 (3) (1982) 527–532.
tween two roughness elements, and this is one reason why the av- [12] S. Shakerin, M. Bohn, R.I. Loehrke, Natural convection in an enclosure with
discrete roughness elements on a vertical heated wall, Int. J. Heat Mass Transf.
erage Nusselt number in the rough case is not much smaller than 31 (7) (1988) 1423–1430.
that in the smooth case. [13] S. Bhardwaj, A. Dalal, S. Pati, Influence of wavy wall and non-uniform heat-
ing on natural convection heat transfer and entropy generation inside porous
complex enclosure, Energy 79 (2015) 467–481.
Declaration of Competing Interest [14] M. Yousaf, S. Usman, Natural convection heat transfer in a square cavity with
sinusoidal roughness elements, Int. J. Heat Mass Transf. 90 (2015) 180–190.
[15] A Shahriari, E.J. Javaran, M. Rahnama, Effect of nanoparticles Brownian motion
The authors declare that they have no known competing finan- and uniform sinusoidal roughness elements on natural convection in an enclo-
cial interests or personal relationships that could have appeared to sure, J. Therm. Anal. Calorim. 131 (3) (2018) 2865–2884.
influence the work reported in this paper. [16] M.M. Kanafi, A.J. Tuononen, Top topography surface roughness power spectrum
for pavement friction evaluation, Tribol. Int. 107 (2017) 240–249.
[17] Y. Xian, P. Zhang, S. Zhai, et al., Re-estimation of thermal contact resistance
Acknowledgement considering near-field thermal radiation effect, Appl. Therm. Eng. 157 (2019)
113601.
[18] J.D. Anderson Jr., Hypersonic and High-Temperature Gas Dynamics, American
This work was supported by Japan Society for the Promotion of Institute of Aeronautics and Astronautics, 2006.
Science (JSPS) KAKENHI [Grant number JP19K14890]. [19] P.L. Roe, Approximate Riemann solvers, parameter vectors, and difference
schemes, J. Comput. Phys. 43 (2) (1981) 357–372.
[20] S. Venkateswaran, C. Merkle, Dual time-stepping and preconditioning for un-
References steady computations, in: 33rd Aerospace Sciences Meeting and Exhibition,
AIAA, Reno, NV, 1995, p. 78.
[1] I. Amber, T.S. O’Donovan, Natural convection induced by the absorption of so- [21] X. Xu, J.S. Lee, R.H. Pletcher, A compressible finite volume formulation for large
lar radiation: a review, Renew. Sustain. Energy Rev. 82 (2018) 3526–3545. eddy simulation of turbulent pipe flows at low Mach number in Cartesian co-
[2] Y. Ma, Z. Yang, Simplified and highly stable thermal Lattice Boltzmann method ordinates, J. Comput. Phys. 203 (1) (2005) 22–48.
simulation of hybrid nanofluid thermal convection at high Rayleigh numbers, [23] R. Ghias, R. Mittal, H. Dong, A sharp interface immersed boundary method for
Phys. Fluids 32 (1) (2020) 012009. compressible viscous flows, J. Comput. Phys. 225 (1) (2007) 528–553.
[3] M.R. Amin, Natural convection heat transfer in enclosures fitted with a peri- [24] M.D. de Tullio, P. De Palma, G. Iaccarino, et al., An immersed boundary method
odic array of hot roughness elements at the bottom, Int. J. Heat Mass Transf. for compressible flows using local grid refinement, J. Comput. Phys. 225 (2)
36 (3) (1993) 755–763. (2007) 2098–2117.

12
B. Ren, C.-G. Li and M. Tsubokura International Journal of Heat and Mass Transfer 173 (2021) 121248

[25] C.G. Li, M. Tsubokura, R. Bale, Framework for simulation of natural convection [29] R.A. Kuyper, T.H. Van Der Meer, C.J. Hoogendoorn, et al., Numerical study of
in practical applications, Int. Commun. Heat Mass Transf. 75 (2016) 52–58. laminar and turbulent natural convection in an inclined square cavity, Int. J.
[26] H. Jia, G. Gogos, Laminar natural convection heat transfer from isothermal Heat Mass Transf. 36 (11) (1993) 2899–2911.
spheres, Int. J. Heat Mass Transf. 39 (8) (1996) 1603–1615. [30] T. Fusegi, J.M. Hyun, Laminar and transitional natural convection in an enclo-
[27] E. Tric, G. Labrosse, M. Betrouni, A first incursion into the 3D structure of nat- sure with complex and realistic conditions, Int. J. Heat Fluid Flow 15 (4) (1994)
ural convection of air in a differentially heated cubic cavity, from accurate nu- 258–268.
merical solutions, Int. J. Heat Mass Transf. 43 (21) (20 0 0) 4043–4056. [33] Hong Kim Kim, Chongam Kim, Accurate, efficient and monotonic numerical
[28] A. Xu, L. Shi, H.D. Xi, Lattice Boltzmann simulations of three-dimensional ther- methods for multi-dimensional compressible flows: Part II: Multi-dimensional
mal convective flows at high Rayleigh number, Int. J. Heat Mass Transf. 140 limiting process, Journal of Computational Physics 208 (2) (2005) 570–615.
(2019) 359–370.

13

You might also like